You are on page 1of 14

J Mater Sci: Mater Electron (2024) 35:728

CO gas‑sensing properties and DFT investigation


of pure and Co‑modified ­MoO3 nanostructures:
effect of solvent composition, deposition time,
and cobalt concentration
G. M. Ramírez1,2, R. Correa1,2, B. García1, Maria de la Luz Olvera3, C. Vargas2, and
T. V. K. Karthik1,*

1
Escuela de Ingeniería y Ciencias, Departamento de Mecánica y Materiales Avanzados, Tecnológico de Monterrey, Avenida Lago de
Guadalupe KM 3.5, Margarita Maza de Juárez, Lopez Mateos, 52926 Mexico City, Mexico
2
Escuela de Ingeniería y Ciencias, Departamento de Ciencias, Tecnológico de Monterrey, Avenida Lago de Guadalupe, KM 3.5,
Margarita Maza de Juárez, Lopez Mateos, 52926 Mexico City, Mexico
3
SEES, Departamento de Ingeniería Eléctrica, Centro de Investigación y de Estudios Avanzados del Instituto Politécnico Nacional,
CINVESTAV-IPN, 07000 Mexico City, Mexico

Received: 29 November 2023 ABSTRACT


Accepted: 30 March 2024 In this study, thin films of pure and cobalt-modified molybdenum oxide ­(MoO3)
were deposited using nebulizing spray pyrolysis (NSP). The research delves into
© The Author(s), under the influence of deposition times (30 and 60 mins), solvents ­(H2O and ­H2O/HCl),
exclusive licence to Springer and cobalt modification (3 %wt and 6 %wt) on the structural, morphological, and
Science+Business Media, LLC, electrical properties of the thin films. XRD and FTIR were employed for structural
part of Springer Nature, 2024 analysis, while SEM and AFM were used to examine morphology and topogra-
phy. X-ray analysis revealed the predominantly amorphous state of most of the
films, and FTIR allowed visualization of different functional groups of ­MoO3
based on deposition time and dopant concentration. In addition, SEM revealed
an increase in grain size with longer deposition times, and AFM demonstrates
that the cobalt-modified films exhibited higher roughness than pure molybde-
num oxide films. Pure ­MoO3 films using ­H2O as a solvent showed the highest
gas-sensing response to CO at 76%, followed by films modified with 6 wt% Co,
which exhibited a sensing response of 66%. The cobalt-modified films exhibited
a lower sensitivity response than pure films, attributed to the formation of cobalt
oxide. The CO adsorption properties on multiple ­MoO3 structures and ­Co3O4
have been optimized and analyzed using density functional theory (DFT). The
best absorption energy of CO onto the different phases of ­MoO3 and ­Co3O4 has
been reported to contrast it with the experimental results. This study marks the

Address correspondence to E-mail: tvkkarthik@tec.mx


E-mail Addresses: a01707978@tec.mx; a01752813@tec.mx; b.garcia@tec.mx; molvera@cinvestav.mx; claudianvh@tec.mx

https://doi.org/10.1007/s10854-024-12501-y

Vol.:(0123456789)
728 Page 2 of 14 J Mater Sci: Mater Electron (2024) 35:728

first report on cobalt-modified molybdenum oxide films, combining theoretical


and practical exploration for CO detection through spray pyrolysis and DFT.

1 Introduction as catalyzers, in field effect transistors, in gas sensors,


and battery electrodes [11].
Currently, the considerable increase in atmospheric Molybdenum oxide has been synthesized using
pollution is caused by deleterious gasses, which con- multiple manufacturing methods categorized into
stitutes a significant challenge for global health [1]. vapor-, solid-, and liquid-phase deposition techniques.
Reports indicate that nearly 3.8 billion individuals Due to its popularity in electronic device fabrication,
suffer from conditions linked to air pollution, and mostly powder or thin films are required [11, 12]. In
approximately 20% of the deaths are attributed to pol- most applications, authors intend to reduce the par-
lution in the domestic environment [2]. CO is posi- ticle size making the material more compelling. The
tioned as one of the polluting gasses at higher risk utilization of thin films in gas-sensing applications
due to its serious toxicity and colorless, tasteless, and offers less sensing material, higher surface area, and
odorless appearance [3, 4]. The human organism has reduction in sensing device size [7]; however, using
an affinity between 250 and 300 times higher for CO powders in gas sensing offers surfaces with very high
compared to oxygen [4]. porosity and abundant oxygen vacancies [11]. It is
The presence of CO in the body causes diseases due important to choose adequate material depending on
to interference with the blood’s ability to transport the device’s characteristics. In addition, for visible-
oxygen, as it combines with hemoglobin in the blood light photocatalysis applications, the thin film has the
forming carboxyhemoglobin [5]. Likewise, it is noted advantage of being reusable, as recovering the cata-
that continuous inhalation of CO over a prolonged lyst after each use requires significantly less effort than
period can trigger cardiovascular diseases, increas- nanoparticles, where catalyst reuse is a laborious pro-
ing the risk of suffering from serious infectious dis- cess [13].
eases [1]. According to the World Health Organization MoO3 thin films have been deposited by various
(WHO), the time-weighted average exposure limit for deposition techniques, such as spray pyrolysis [7–9],
CO is 9 ppm for 8 h and 26 ppm for 1 h [6]. Conse- chemical vapor deposition [14, 15], sputtering [16],
quently, the creation of highly sensitive and selective gel processing [17], and thermal evaporation [18].
sensors capable of detecting concentrations below The ultrasonic spray pyrolysis technique has attracted
the threshold limit of CO gas is crucial. This will help immense attention to deposit high-quality thin films
reduce the potential risk to human health and ensure because of its one-step deposition. In addition, it is
safety in the public sphere. a simple, fast, and cost-effective (equipment and raw
Transition metal oxides semiconductors (MOS) have materials) technique relative to several other bottom-
recently received much attention due to their diverse up synthesis methods [14, 15, 18]. Response and recov-
structural, morphological, and electrical properties [7]: ery times are the most critical factors for a gas sensor
Among different MOS, molybdenum oxide ­(MoO3) is to be utilized in real-time applications. These prop-
a n-type semiconductor that has an energy bandgap erties can be improved by increasing gas molecules
of ~ 3.20 eV, excellent electron mobility and economi- adsorption and desorption rates by applying ther-
cal phase costs, controlled production, non-toxicity, mal energy or temperature. However, increasing the
and multiple valence states [8, 9]. The three different temperature of a gas sensor results in higher power
crystalline structures are thermodynamically stable consumption. Doping the semiconductor oxide with
orthorhombic ­MoO3 (α-MoO3), metastable monoclinic metals is known to be a very suitable method for opti-
­MoO3 (β-MoO3), and hexagonal ­MoO3 (h-MoO3) [10]. mizing the parameters of metal oxide gas sensors [7–9,
The interesting structural and electrical properties of 19]. This is because surface metallic dopants cause (a)
­MoO3 such as crystalline structure, surface resistance, modification of gas molecule adsorption sites or (b)
surface oxygen vacancies, and physicochemical and pinning of the Fermi energy at the surface [20].
mechanical characteristics, such as refractive index, For example, Comini et al. [21] studied sol–gel (SG)
energy gap, and hardness out of ß-MoO3 [8] make it and radio-frequency (RF) sputtering techniques for
suitable for various technological applications such obtaining ­MoO3 thin films. RF-deposited films have
J Mater Sci: Mater Electron (2024) 35:728 Page 3 of 14 728

a well-developed nanoparticle microstructure and by NSP. Finally, in this work, a gas-sensing mecha-
a larger surface area than SG-deposited films. The nism of pure and Co-modified ­MoO3 sensors for CO
results have revealed that the ­MoO3 thin films pre- has been reported and discussed in detail.
pared by these techniques are highly sensitive to CO
with very fast response and recovery in the order of a
minute. In addition, Gouider Trabelsi et al. [7] used a
nebulizing spray pyrolysis deposition method to fab- 2 Experimental details
ricate highly sensitive Co-doped ­MoO3 films to detect
ammonia gas. The authors mentioned that the films 2.1 Film preparation
with 3% Co-doped ­MoO3 exhibited faster response
and recovery times than sensors with undoped ­MoO3 Thin films of ­MoO3 were deposited on thoroughly
film, which is critical for ­MoO3 sensor practical appli- cleaned and dried glass substrates. Two precursor
cations. Furthermore, the ammonia selectivity of the solutions (Solution A and Solution B) were prepared,
doped ­MoO3 sensors was found to be very remark- Solution A was made by mixing 0.04 M ammonium
able. In addition, Yoon Ho Cho et al. [19] achieved heptamolybdate tetrahydrate ((NH4)6 ­Mo7O24·4H2O)
an ultraselective and ultrasensitive trimethylamine in deionized water, and Solution B was prepared by
(TMA) sensor using ­MoO3 nanoplates prepared by mixing ­H2O and HCl as solvents in a 3:1 ratio, respec-
ultrasonic spray pyrolysis. The nanoplates showed tively. The total volume of each solution was 100 ml,
an unusually high response to 5 ppm TMA (ratio of and their pH values were 5 and 1 for Solution A and
resistance to air and gas = 373.74) at 300 °C with a Solution B, respectively. Appropriate quantities of
detection limit as low as 45 ppb. cobalt chloride hexahydrate ­(CoCl2·6H2O) were added
In addition, the effect of deposition time [7], the to Solution B for modified this film. Two different
effect of calcination [10, 19], and different dopants over weight ratios of [Co]/[Mo]: 3% and 6% were utilized
the gas-sensing properties have also been studied for to observe the effect of Co concentration. All the films
­MoO3 films [7, 22]. However, a study of the effect of were deposited at 200 °C in a tin bath, with a 1 ml/min
the solvent composition, such as pure ­H2O or ­H2O/ flow rate at 30 and 60 mins utilizing NSP. After the
HCl, on the films yet to be reported. We believe that deposition, films were allowed to cool to room tem-
the change in the solvent composition affects the solu- perature and were calcined at 350 °C for 2 h.
bility of the precursor, which in turn changes the pH
of the solution and the thin film growth. Therefore, in
this work, we have deposited molybdenum oxide thin 2.2 Characterization techniques
films by NSP with different Co-dopant concentrations,
solvent compositions, and at different deposition XRD analysis was performed using a PANalytical
times. The changes in the structural, topographical, diffractometer with CuKα at 20 mA and 40 KV was
morphological, and electrical or gas-sensing properties carried out to identify the phase compound and the
of ­MoO3 nanostructures were analyzed. crystalline structure. Scanning electron microscopy
The density functional theory (DFT) calculation is (SEM JEOL JSM-6360 LV) was used to analyze the
a technique for analyzing electronic, energetic, and morphology and thickness of the prepared films. The
adsorption properties at the atomic level. The motiva- topography and surface roughness of the synthesized
tion of this study is to establish a correlation between thin films were examined by atomic force microscopy
the CO adsorption properties on the amorphous (A), (AFM-Park Systems model XE7). The characteristic
hexagonal (h), and alpha ( 𝛼 ) phases of ­MoO3 and the functional groups were obtained by Fourier Transform
experimental results. First principles’ methods based Infrared Spectroscopy (FTIR-Shimadzu Affinity-1S).
on DFT have been used to study the sensing mecha- Silver contacts were painted over sensor surfaces for
nism of gas molecules adsorbed on a ­MoO3 surface the gas-sensing measurement and let dry for 20 min at
[23]. However, the sensing mechanism of a ­Co3O4 room temperature. The home-made gas-sensing sys-
structure is considered in this work due to the experi- tem reported in our previous work [24] was utilized to
mental Co modification of ­MoO3 thin films. obtain all sensors’ sensing response (SR). SR is defined
We are the first to report the CO gas-sensing proper- in Eq. (1), where Ro is the resistance in vacuum and Rg
ties of pure and Co-modified ­MoO3 thin films obtained is the resistance in gas.
728 Page 4 of 14 J Mater Sci: Mater Electron (2024) 35:728

RO − Rg the mentioned energies in the equation relate to the


SR = × 100. (1) equivalently relaxed minimum energy structures.
Ro

2.3 Density functional theory simulation 3 Results and discussion

Geometry optimization and adsorption energy are 3.1 E


 ffect of solvent composition based
obtained by DFT techniques using the ORCA pro- on deposition time
gram package (5.0.3) [25]. Geometry optimizations
are treated using Becke’s three-parameter hybrid func- 3.1.1 FTIR‑DRX analysis
tional with the Lee–Yang–Parr correlation functional
(B3LYP) [26] in combination with LANL2DZ basis set Figure 1 shows the FTIR spectra of ­MoO3 thin films
on unit cells of multiple ­MoO3 and ­Co3O4 structures. at different deposition times using ­H2O and ­H2O/
Cobalt oxide structure is considered instead of pure HCl (Fig. 1a, b) as solvents. The ­H2O spectra at 30
Co due to the experimental results obtained (refer to and 60 min of deposition showed absorption bands
Sect. 3.2.1: FTIR-XRD analysis). For CO-MoO3/Co3O4 at 761.65 ­cm−1 that are assigned to the Mo–O stretch-
interaction, B3LYP with 6-31G* and LANL2DZ level of ing vibration and the bands at 885.75 and 908.31 ­cm−1
theory was used [27]. The adsorption energies ( Eads ) associated with the Mo=O stretching vibration. On the
between CO and ­M oO 3 and ­C o 3O 4 structures are other hand, samples with ­H2O/HCl solvents depos-
defined in Eqs. (2) and (3): ited at 30 and 60 min of deposition showed the bands
( ) at 761.65 ­cm−1 associated with the Mo–O stretching
Eads = EMoO3 /CO − EMoO3 + ECO , (2) vibration and the bands at 887.25 and 1072.68 ­cm−1 that
are associated with the Mo=O stretching vibration.
( ) From the FTIR spectra of Fig. 1, it can be confirmed
Eads = ECo3 O4 /CO − ECo3 O4 + ECO , (3) the formation or presence of ­MoO3. Sudesh Kumari
et al. [12] have reported similar characteristic absorp-
where EMoO3 /CO corresponds to the CO adsorption tion bands of molybdenum oxide in the wavenumber
energy in the different crystal structures of ­MoO3, range of 400–1200 ­cm−1.
ECo3 O4 /CO corresponds to the CO adsorption energy The difference in peak intensity in Fig. 1a may be
in ­Co3O4 structure, EMoO3 corresponds to the ­MoO3 attributed to variations in the concentration of the
adsorption energy, ECo3 O4 corresponds to the ­Co3O4 functional group or changes in the molecular envi-
adsorption energy structure, and then ECO corre- ronment, an increase in the deposition time resulted
sponds to the energy of an isolated CO molecule. All in a red shift in the Mo=O stretching vibration, which

Fig. 1  FTIR spectra of ­MoO3 thin films at 30 and 60 min of deposition a pure with H
­ 2O as solvent and b pure with ­H2O/HCl as solvents
J Mater Sci: Mater Electron (2024) 35:728 Page 5 of 14 728

corresponds to a change in the hybridization state or films of 100, 200, and 300 nm thickness. As mentioned
electron distribution in the molecular bond. In Fig. 1b, in the experimental details, the substrates were placed
the longer the deposition time, the greater an increase in a tin bath during the deposition which resulted in
in transmittance is observed, which resulted due to certain diffraction peaks at 2θ = 32° and 67° of this
increased crystallinity of the material. To confirm the material.
crystalline structure of the thin films, XRD analysis The peaks at (210) and (320) correspond to the hex-
of the samples was performed and reported in Fig. 2. agonal phase of ­MoO3 according to Zhang et al. [28].
Figure 2 shows the X-ray diffraction analysis of The diffraction peaks coincide with JCPDS card No.
molybdenum oxide thin films at different deposition 21-0569 (space group P63), indicating that hexagonal
times using ­H2O/HCl as a solvent, revealing to the ­MoO3 was obtained. According to Caique et al. [29]
naked eye an amorphous or nanostructured nature and Amira et al. [7], the peak at 2θ = 26.5° corresponds
due to the presence of a wide hump. Sudesh Kumari to the alpha phase of ­MoO3. Furthermore, a 30 min
et al. [12] also obtained similar results for ­MoO3 thin deposition time led to a (210) peak intensity increase,
indicating an anisotropic growth [29] of the h-MoO3
structure and a higher crystallinity. Therefore, two
phases of hexagonal and alpha ­MoO3 were obtained.

3.1.2 SEM‑AFM analysis

With FTIR, it was possible to verify the presence of


­MoO3, however, to achieve a better understanding
of the effect of deposition time, the morphological
and topographic analysis shown in Fig. 3 was carried
out. In the SEM images (Fig. 3a, b), a smooth surface
with homogeneously distributed nanoparticles and
some agglomerates of micron size of 3.33 and 4.8 µm
respectively can be observed. When adding acid to the
solvent (Fig. 3c, d), two different morphologies were
observed: 1. Microcrystal formation is exhibited due
to the coalescence effect, and 2. Microparticles are uni-
Fig. 2  X-ray diffraction patterns of ­MoO3 at 30 and 60 mins of formly distributed on the surface. Authors believe that
deposition with ­H2O/HCl as a solvent the second morphology possesses greater roughness

Fig. 3  SEM images and 3D AFM images of pure with ­H2O as solvent (a, e and b, f) and pure with ­H2O/HCl as solvents (c, g and d, h)
­MoO3 thin films deposited at 30 and 60 mins, respectively
728 Page 6 of 14 J Mater Sci: Mater Electron (2024) 35:728

and is favorable for sensing. Figure 3c shows the for-


mation of cubic microcrystals with a size of approxi-
mately 2.5–3 µm and Fig. 3d exhibits dendritic struc-
tures with sizes of 5–7 µm. With the transverse SEM
images (Fig. S1), it can be confirmed that the longer
the deposition time, the thickness increases from 224
nm to 88.4 µm and the particle size because there is a
greater amount of material deposited.
It is important to observe that the addition of HCl
in the solvent resulted in different surface morpholo-
gies. Reduction of pH by addition of HCl reduced the
particle size from 4.8 to 3 µm, which is a general phe-
nomenon reported by several authors [30–33] that at
lower pH, formation of ­H3O+ ions is formed, which
prevents the hydrolysis of ­MoO3 and results in smaller
aggregates. On the other hand, authors believe that
addition of HCl could resulted in different cations Fig. 4  FTIR spectra of ­MoO3 thin films of different deposition
of Mo like ­Mo4+ o ­Mo3+ in the spray solution, which times and different percentages of Co
could have resulted in the formation of microcrystals
observed for samples with HCl as solvent [30]. How-
ever, further studies are necessary to corroborate the
result in detail.
During the AFM measurements (Fig. 3e–h), the
fewer rough areas were analyzed, for equipment
safety reasons. In all the films, a very homogeneous
topography is observed. Roughness was calculated
from the AFM height profile of the 5 × 5 µm scanned
area. The 30 and 60 mins ­H2O solvent films (Fig. 3e, f)
showed very similar roughness values of 4.5 and 4.9
nm, respectively, while the roughness values obtained
for the ­H2O/HCl of 30 and 60 min (Fig. 3g, h) were 8.8
and 15.4 nm, respectively. Likewise, it can be observed
that the longer the deposition time, in both cases,
the thickness of the film increases. These substantial
changes in roughness with changes in thickness values
can be attributed to the reflective process of nuclea- Fig. 5  X-ray diffraction patterns of ­MoO3 thin films of different
tion, coalescence, and continuous film growth [34]. deposition times and different percentages of Co

3.2 Effect of Co modification are characteristic of C–O [36]. It can be confirmed from
the FTIR spectra that Co has formed ­Co3O4 instead
3.2.1 FTIR‑DRX analysis of incorporating it into the ­MoO3 crystal structure.
The time effect in both cases with 3% and 6% of Co
Figure 4 shows the FTIR spectra of ­MoO3 thin films at
produced an increase in transmittance, correspond-
different deposition times and different cobalt percent-
ing to the improved crystallinity of the sample with
ages. The spectra showed absorption bands at 625.29
increased deposition time. Besides that, no significant
and 639.30 ­cm−1, which are the characteristics of Co–O
changes between different Co samples percentages are
[35], and the band at 639.30 ­cm−1 associated with the
observed in spectra.
stretching vibration of Mo–O. The bands at 878.31 and
Figure 5 shows the X-ray diffraction analysis of
904.58 ­cm−1 are associated with the Mo=O stretching
molybdenum oxide thin films at different deposition
vibration and the bands at 1060.42 and 1105.29 ­cm−1
J Mater Sci: Mater Electron (2024) 35:728 Page 7 of 14 728

times and different percentages of Cobalt, revealing Co (Fig. 6e, f) obtained a roughness of 46.6 and 20.8
with the naked eye an amorphous or nanostructured nm, respectively. Likewise, it can be observed that the
nature due to the presence of a wide hump. longer the deposition time, in all cases, the thickness of
The peak at 2θ = 19° confirms the formation of the film increases. These substantial changes in rough-
­Co3O4 according to Bachiri et al. [37]. The peaks at 2θ ness with changes in thickness values can be attributed
= 26°, 30°, and 43.5° correspond to the hexagonal phase to the reflective process of nucleation, coalescence, and
of ­MoO3 according to Zhang et al. [28]; the peak at 2θ continuous film growth [34]. The topography with Co
= 26° and 39° corresponds to the alpha phase of ­MoO3 modification presented greater roughness than the
according to Caique et al. [29] and Amira et al. [7]; and pure films, which would favor sensing.
the peak at 2θ = 32° refers to tin due to the methodol-
ogy used for the deposition of the films.
Therefore, with this image, it can be seen that two
3.2.3 SEM analysis
phases of ­MoO3 were obtained, hexagonal and alpha,
with the hexagonal being the most prominent. Fur-
In the SEM images (Fig. 7a–f), a smooth surface with
thermore, the films with the longest deposition time
homogeneously distributed microparticles can be
(60 min) presented a more crystalline structure than
observed, and they exhibit a type of morphology
the films with a shorter deposition time.
with the formation of crystals or microstructures due
to the coalescence effect. Figure 7a shows the forma-
3.2.2 AFM analysis tion of cubic crystals with a size of about 2.5–3 µm,
and Fig. 5b exhibits dendritic structures with sizes of
During the AFM measurements (Fig. 6a–f), the less 5–7 µm.
rough areas were analyzed, for equipment safety rea- Films with 3% of Co (Fig. 7c, d) show preferen-
sons. A very homogeneous topography is observed in tial growth with structures of a length of 5 µm and
all the films. Roughness was calculated from the AFM particles of less than 1 µm while films with 6% of
height profile of the 5 × 5 µm scanned area. The 30 and Co (Fig. 7e, f) have anisotropic growth with struc-
60 mins films (Fig. 6a, b) showed roughness values of tures with a length of 3 µm and particles of 1 µm to
8.8 and 15.4 nm, respectively. 2 µm. With the SEM images, it can be confirmed that
The 30- and 60-min films with 3 wt% Co (Fig. 6c, the longer the deposition time, the larger the parti-
d) obtained a roughness of 26.9 and 56.6 nm, respec- cle size because there is a greater amount of material
tively. Finally, the 30- and 60-min films with 6 wt% deposited, and agglomeration of this occurs and with

Fig. 6  3D AFM images of pure (a, b), 3 wt% Co-doped, (c, d) and 6 wt% Co-doped, (e, f) ­MoO3 thin films deposited at 30 and 60 mins,
respectively
728 Page 8 of 14 J Mater Sci: Mater Electron (2024) 35:728

Fig. 7  SEM images of


­MoO3 of pure (a, b), 3 wt%
Co-doped (c, d), and 6 wt%
Co-doped (e, f) ­MoO3 thin
films deposited at 30 and 60
mins, respectively

Co modification a different, rougher morphology is from the figures that majorly the responses are lin-
observed, ideal for sensing. ear with respect to gas concentration. However, it is
important to note that for some samples, a saturation
of response is noticed after 200 ppm (See Fig. 9a). It
3.3 Gas sensing
is well known that the sensing performance depends
The CO-sensing properties of ­MoO3 thin films were mainly on the surface oxygen vacancies, which acts
analyzed at different gas concentrations from 10 to as the adsorption sites and further gives the sensing
500 ppm at 150 °C, 200 °C, and 300 °C. In general, signal. Essentially, the oxygen-deficient sites, which
the gas-sensing properties were influenced by sev- favors the gas-sensing mechanisms, are closely tied
eral parameters such as morphology, oxygen vacan- to its uniformity in the films [1]. Therefore, as long
cies, surface roughness, and crystalline structure. The as the surface contains abundant surface adsorbed
above-mentioned parameters can be optimized by oxygen, the sensing response shows a linear behav-
varying sensor-operating temperature, Co concentra- ior with respect to gas concentration after which
tions, deposition time, and the solvent used. the response saturates, which is generally termed
Figures 8 and 9 show the ­S R curves of the sen- as detection limit. Authors believe that the linear
sors to CO at different operation temperatures of all behavior is observed between CO concentration
the sensors obtained in this work. It can be observed and S R which is due to an increase in the number
J Mater Sci: Mater Electron (2024) 35:728 Page 9 of 14 728

Fig. 8  Sensing response of the sensors to CO as a function of gas concentration at different temperatures, pure with H
­ 2O as solvent (a,
b) and pure with ­H2O/HCl as solvents (c, d) ­MoO3 thin films deposited at 30 and 60 mins, respectively

of interactions between test gas molecules and the respectively. A slightly higher sensitivity can be
semiconductor surface. observed for the sensors prepared with ­H 2O. On
It is well known that the gas-sensing responses the contrary, the characterization by SEM and AFM
directly depend on the sensor operation temperature. demonstrated that the films with ­H2O/HCl exhibit
At lower temperature (below 100 °C), only physisorp- a relatively homogeneous morphology. Further
tion is observed and no electron interchange occurs experiments are necessary for concluding the higher
between test gas and sample surface. At tempera- responses of sensors with ­H2O as solvent.
tures between 100 and 300 °C, the chemisorption of Figure 9a–d shows the response of the sensors
atmospheric oxygen occurs resulting higher sens- with 3 wt% and 6 wt% Co-modified ­MoO3 thin films,
ing responses. In this work, sensing responses were respectively. 6 wt% modified films show a very high
observed only at 150 and 200 °C. At higher tempera- SR, and films deposited for 60 min exhibit a better
tures (above 200 °C), the sensor surface obtained in structure, and higher SR. Finally, it is also evident
this work undergoes an oxidation process because the that films deposited for 60 min show higher SR than
samples are amorphous (refer to Sect. 3.2.1: FTIR-DRX films deposited at 30 min, irrespective of solvent and
analysis). Co concentration. From XRD and FTIR analysis, it
Figure 8a–d shows the response of the sen- is confirmed that the films with a deposition time
sors prepared with ­H2O and ­H2O/HCl as solvents, of 60 min are more crystalline which subsequently
increased the response to CO.
728 Page 10 of 14 J Mater Sci: Mater Electron (2024) 35:728

Fig. 9  Sensing response of the sensors to CO as a function of gas concentration at different temperatures, 3 wt% Co-doped, (a, b) and 6
wt% Co-doped (c, d) ­MoO3 thin films deposited at 30 and 60 mins, respectively

As in XRD analysis, it is evident the formation of repeatability tests are necessary to confirm the SR, 66%
cobalt oxide, which is a p-type semiconductor; the of SR was observed for the one prepared with only
p–n junctions between cobalt oxide and molybdenum ­H2O/HCl as solvent. On the other hand, SR around
oxide resulted in a decrease in the SR for the modified 70% and 57% was recorded for 6 wt% and 3 wt% mod-
thin films. ified thin films, respectively. The authors believe that
For comparative purposes, SR obtained for all sen- the decrease in the SR for the modified samples is also
sors deposited at 30 and 60 min at 300 °C are plotted in due to a slight increase in the particle size observed
Fig. 10a and b, respectively. From Fig. 10a, it is evident from SEM analysis. Tayier Yunusi et. al. [31] also
that the films with ­H2O/HCl solvent show the highest obtained ethanol gas detection result for ­MoO3 thin
response of 76% at 500 ppm; however, it is important films with a very similar SR of 64.1 and for acetone gas
to mention that further repeatability tests are neces- with a lower SR of 36.7.
sary to confirm the SR, 54% of SR was observed for the
one prepared with only ­H2O as solvent. On the other 3.4 Density functional theory simulation
hand, SR around 67% and 44% was recorded for 6 wt%
and 3 wt% modified thin films, respectively. To identify the absorption behavior of CO gas, the
From Fig. 10b, it is evident that the films with first-principles calculations based on DFT study
­H2O solvent show the highest response of 74% at 500 considering different positions configurations of
ppm; however, it is important to mention that further CO molecules were performed to compare the
J Mater Sci: Mater Electron (2024) 35:728 Page 11 of 14 728

Fig. 10  Sensing response of the sensors to CO as a function of the gas concentration at 350 °C with different percentages of cobalt, sol-
vents, and deposition times (a, b) at 30 and 60 mins, respectively

adsorption energies towards the hexagonal (h), alpha The optimized adsorption models of (h), ( 𝛼 ), (A)
( 𝛼 ), and amorphous (A) ­MoO3 structures. Moreover, and ­C o 3O 4 representative molecules are shown in
adsorption energy from ­Co3O4 was also calculated. Fig. 11. Here, the red, pale blue, pale pink, and gray

Fig. 11  Four optimized complex structures: A-MoO3 (a), α-MoO3 (b), h-MoO3 (c), and ­Co3O4 (d)
728 Page 12 of 14 J Mater Sci: Mater Electron (2024) 35:728

balls were O, Mo, Co, and C atoms, respectively. In Even though the samples worked for CO gas sens-
Table 1, it lists the specific simulation results of all ing, more crystalline samples are expected to show
adsorption systems. It was reported that the greater a better sensing response, considering the adsorption
the adsorption energy is, the stronger the adsorption energies obtained during the simulation, and the nega-
capacity of the structure for the CO gas molecule is tive values obtained that refer to an oxidation process
[22]. that can lead to a more crystalline structure formation
The calculation results exhibit that the ­Co3O4 had on ­MoO3 films. It must be highlighted that interactions
the most negative adsorption for carbon monoxide between CO and Co-modified ­MoO3 films may change
(− 2.65 eV), given that a ­MoO3 molecule exhibited less when having a more crystalline structure, having a
adsorption energy (− 0.82 eV). The greater adsorp- direct impact on the sensing response system. Table 2
tion energy ( Eads ) was calculated to be between CO shows a comparison of results obtained in this work
and h-MoO3 (− 12.86 eV). It is noted that the amor- with previously reported ­MoO3 sensors. To the best of
phous structure presented the weaker adsorption tour knowledge, we are the first to report ­MoO3 sen-
for CO, which explains the sensing response value sors obtained by NSP for detecting CO.
since the amorphous structure is the highest in films.
However, due to the Co-modified ­MoO3, the ­Co3O4
structure formed, and its adsorption energy, authors 4 Conclusions
believe that CO firstly interacts with it given that it is
a p-type. Thus, it is suggested that ­Co3O4 tends to fill In this study, NSP successfully deposited thin films
its vacancies while interacting with CO molecules, of pure and cobalt-modified ­M oO 3. The research
and once full, the sensing mechanisms begin. This focused on systematically examining the impact
explains the ­SR decreases when having a doped film of deposition times, solvents, and cobalt modifica-
n + p-type rather than having a pure one (n-type). tion on the structural, morphological, and electrical
From analyzing Figs. 10, 11, and Table 1, it can be properties of thin films. Structural analysis by XRD
discussed that the more crystalline the ­M oO 3 film revealed the predominantly amorphous state of most
is, the greater the absorption energy and sensing of the films, while FTIR showed different functional
response. As it is shown in Table 1, A-MoO 3 pre- groups depending on the deposition time and cobalt
sents the lowest negative adsorption energy (− 9.60 concentration. Morphological and topographic exam-
eV) due to its arbitrary molecules’ position, while inations using SEM and AFM indicated an increase
h-MoO 3 the highest. Thus, the S R obtained can be in grain size with deposition times of 60 min and
justified considering that it is shown in XRD spectra greater roughness in films modified with cobalt; spe-
that there is more presence of ­MoO3 amorphous state cifically the films with 6%wt Co exhibited the greatest
than the crystalline ones. roughness.
The gas-sensing properties were influenced by sev-
eral parameters such as morphology, oxygen vacan-
Table 1  Adsorption energies cies, surface roughness, and crystal structure. It could
Structure Eads (eV)
from CO complex systems be observed that the detection response of the sensors
MoO3 (A) − 9.60 decreases as the temperature increases. The authors
MoO3 (h) − 14.37 believe that the sensors undergo an oxidation process
MoO3 (𝛼) − 12.86 with an increase in temperature because the samples
Co3O4 − 2.65 are amorphous, the results at 300 °C showed that pure

Table 2  A comparison of


Synthesis method Gas Response (%) Temp. (°C) Conc. (ppm) References
­MoO3 based gas sensors
NSP CO 76 300 500 This work
Grinding + sonication Alcohol 33 300 100 [38]
Spray pyrolysis TMA 12 30 0.5 [39]
Chemical bath deposition NO2 71.8 200 100 [40]
Hydrothermal Acetona 38 260 200 [41]
J Mater Sci: Mater Electron (2024) 35:728 Page 13 of 14 728

­MoO3 films using ­H2O as a solvent exhibited the high- Funding


est SR, followed by the modified films with 6 wt% Co.
Analysis of the cobalt-modified films suggested lower Authors did not receive any special funding from any
sensitivity compared to the pure films, attributed to project. Therefore no funding information is added.
the formation of cobalt oxide rather than forming as
a dopant.
DFT calculations were done to add a better under- Data availability
standing on the interactions between CO and the com-
plex systems of the films, showing a greater adsorp- We state that our manuscript is original and unpub-
tion energy with ­Co3O4 than ­MoO3, explaining the lished, and all data are true in the original manuscript
sensing response decrease. Also, multiple ­MoO3 crys- and is not considered for publication elsewhere. We
talline phases were studied, evidencing a greater CO would be grateful if our paper could be reviewed and
interaction when having a crystalline system. Taking considered for publication in the Journal of Materials
all into consideration, this research provided valuable Science: Materials in Electronics.
information on the potential applications of cobalt-
modified ­MoO3 films and several flaws that need to
be improved in future works in order to lay greater Declarations
studies in gas sensing.
Conflict of interest The authors declare that they
have no known competing financial interests or per-
Acknowledgements sonal relationships that could have appeared to influ-
ence the work reported in this paper.
We thank Javier Ulises Arreola for providing the space
and precursors for chemical synthesis, and Mario
Supplementary Information The online version
Andrei Loperena for providing the space for spray
contains supplementary material available at https://​
pyrolysis. Subsequently, we thank the help provided
doi.​org/​10.​1007/​s10854-​024-​12501-y.
by Dr. Dulce Viridiana Melo for the AFM measure-
ments, we thank Juan Jesús Rocha and Kevin Rueda
for the SEM measurements and the technical assis- References
tance for sensing by Miguel Ángel Luna. Finally, we
thank Tecnológico de Monterrey for the socioeconomic 1. A. Ani, P. Poornesh, A. Antony, S. Chattopadhyay, Sens.
scholarships for our studies. Actuators B 399, 134827 (2024)
2. S.M. Majhi, A. Mirzaei, H.W. Kim, S.S. Kim, T.W. Kim,
Nano Energy 79, 105369 (2021)
Author contributions 3. Z.A. Ansari, S.G. Ansari, T. Ko, J.H. Oh, Sens. Actuators
B 87, 105 (2002)
All authors contributed to the conception and design
4. S. Shrivastava, D. Mahana, S. Nehra, S. Gangwar, S. Singh,
of the study. G. M. Ramírez contributed to the experi-
C.S. Yadav, S.K. Muthusamy, A. Dogra, Sens. Actuators B
mentation, analysis, and first draft of the manuscript.
400, 134882 (2024)
R. Correa contributed to the simulations, analysis,
5. D. Zhang, C. Jiang, J. Liu, Y. Cao, Sens. Actuators B 247,
and first draft of simulations. B. García contributed
875 (2017)
to AFM analysis, English revision, and finalization of
6. S. Ghosh, M. Narjinary, A. Sen, R. Bandyopadhyay, S. Roy,
the manuscript. M. de la L. Olvera contributed to the
Sens. Actuators B 203, 490 (2014)
design of experiments and analysis of the detection
7. A. Ben Gouider Trabelsi, F.H. Alkallas, M. Aslam Man-
results. C. Vargas contributed to the design of experi-
thrammel, M. Shkir, S. AlFaify, Results Phys. 43, 106036
ments and the interpretation of simulation results.
(2022)
T. V. K. Karthik contributed to the overall design of
8. R. Jansi, M.S. Revathy, S. Vinoth, A. Kumar, R.S.R. Isaac,
the research work and analysis and interpretation of
N. Deepa, A.M. Al-Enizi, M. Ubaidullah, B. Pandit, M.
XRD, FTIR, SEM, and finalization of the manuscript.
Shahazad, M. Gupta, Opt. Mater. (Amst.) 145, 114464
All authors read and approved the final manuscript.
(2023)
728 Page 14 of 14 J Mater Sci: Mater Electron (2024) 35:728

9. V. Ganesh, I.S. Yahia, Opt. Mater. (Amst.) 144, 114480 29. C.D.A. Lima, J.V.B. Moura, G.S. Pinheiro, J.F.D.F. Araujo,
(2023) S.B.S. Gusmão, B.C. Viana, P.T.C. Freire, C. Luz-Lima,
10. S. Kumari, P. Singh, H. Singh, K. Singh, A. Kumar, S. Ceram. Int. 47, 27778 (2021)
Kumar, A. Thakur, J. Mater. Sci. 32, 24990–24996 (2021) 30. B. Yusuf, M.R. Hashim, M.M. Halim, Results Phys. 45,
11. T.H. Chiang, H.C. Yeh, Materials 6, 4609 (2013) 106229 (2023)
12. S. Kumari, K. Singh, P. Singh, S. Kumar, A. Thakur, 31. T. Yunusi, C. Yang, W. Cai, F. Xiao, J. Wang, X. Su,
(123AD). Ceram. Int. 39, 3435 (2013)
13. M. Sarma, M. KumariJaiswal, S. Podder, J. Bora, S. Kar- 32. X. Fu, P. Yang, X. Xiao, D. Zhou, R. Huang, X. Zhang, F.
makar, B. Choudhury, A. RatanPal, Physica B 670, 415354 Cao, J. Xiong, Y. Hu, Y. Tu, Y. Zou, Z. Wang, H. Gu, J.
(2023) Alloys Compd. 797, 666 (2019)
14. A. Abdellaoui, L. Martin, A. Donnadieu, Physica Status 33. L. Khandare, S.S. Terdale, D.J. Late, Adv. Device Mater.
Solidi (a) 109, 455 (1988) 2, 15 (2016)
15. T. Ivanova, M. Surtchev, K. Gesheva, Mater. Lett. 53, 250 34. I. Dundar, M. Krichevskaya, A. Katerski, I.O. Acik, R. Soc.
(2002) Open Sci. 6, 181578 (2019)
16. S. Uthanna, V. Nirupama, J.F. Pierson, Appl. Surf. Sci. 256, 35. F. Zhang, C. Yuan, X. Lu, L. Zhang, Q. Che, X. Zhang, J.
3133 (2010) Power Sources 203, 250 (2012)
17. M. Dhanasankar, K.K. Purushothaman, G. Muralidharan, 36. A. Stadnik, F.Q. Mariani, F.J. Anaissi, S. Afr. J. Chem. 70,
Appl. Surf. Sci. 257, 2074 (2011) 137 (2017)
18. B. Han, M. Gao, Y. Wan, Y. Li, W. Song, Z. Ma, Mater. Sci. 37. A. El Bachiri, L. Soussi, O. Karzazi, A. Louardi, A. Rmili,
Semicond. Process. 75, 166 (2018) H. Erguig, B. El Idrissi, Spectrosc. Lett. 52, 66 (2019)
19. Y.H. Cho, Y.N. Ko, Y.C. Kang, I.D. Kim, J.H. Lee, Sens. 38. F. Ji, X. Ren, X. Zheng, Y. Liu, L. Pang, J. Jiang, S. Liu,
Actuators B 195, 189 (2014) Nanoscale 8, 8696 (2016)
20. A. Galdikas, A. Mironas, A. Šetkus, Sens. Actuators B 26, 39. R. Pandeeswari, B.G. Jeyaprakash, Biosens. Bioelectron.
29 (1995) 53, 182 (2014)
21. E. Comini, G. Faglia, G. Sberveglieri, C. Cantalini, M. 40. A.A. Felix, R.A. Silva, M.O. Orlandi, CrystEngComm 22,
Passacantando, S. Santucci, Y. Li, W. Wlodarski, W. Qu, 4640 (2020)
Sens. Actuators B 68, 168 (2000) 41. H. Yan, P. Song, S. Zhang, Z. Yang, Q. Wang, RSC Adv. 5,
22. Y. Tutel, M.B. Durukan, S.O. Hacioglu, U.C. Baskose, 72728 (2015)
L. Toppare, H.E. Unalan, Appl. Mater. Today 35, 101924
(2023) Publisher’s Note Springer Nature remains neutral with
23. K. Xu, N. Liao, B. Zheng, H. Zhou, Phys. Lett. A 384, regard to jurisdictional claims in published maps and
126533 (2020) institutional affiliations.
24. R. Lozano-Rosas, J.M. Bravo-Arredondo, V.K. Karthik-
Springer Nature or its licensor (e.g. a society or other partner)
Tangirala, M.J. Robles-Águila, Appl. Phys. A 129, 1 (2023)
holds exclusive rights to this article under a publishing
25. ORCA: An Ab Initio, DFT and Semiempirical Electronic
agreement with the author(s) or other rightsholder(s);
Structure Package (Version 5.0.3). Max Planck Institute for
author self-archiving of the accepted manuscript version of
Chemical Energy Conversion. (2022).
this article is solely governed by the terms of such publishing
26. A.D. Becke, Phys. Rev. A (Coll Park) 38, 3098 (1988)
agreement and applicable law.
27. A.S. Rad, S.S. Shabestari, S. Mohseni, S.A. Aghouzi, J.
Solid State Chem. 237, 204 (2016)
28. C.C. Zhang, L. Zheng, Z.M. Zhang, R.C. Dai, Z.P. Wang,
J.W. Zhang, Z.J. Ding, Phys. Status Solidi B 248, 1119
(2011)

You might also like